You are on page 1of 17

Resources, Conservation and Recycling 50 (2007) 442458

A process model to estimate the cost of industrial


scale biodiesel production from waste cooking
oil by supercritical transesterification
J.M.N. van Kasteren a, , A.P. Nisworo b
aTelos, Brabants Institute for Sustainable Development, P.O. Box 90153,
5000 LE Tilburg, The Netherlands
b Eindhoven University of Technology, Department of Chemical Engineering, Process and Product
Design, Den Dolech 2, 5612 AX Eindhoven, The Netherlands

Received 17 November 2005; received in revised form 30 June 2006; accepted 5 July 2006
Available online 22 August 2006

Abstract

This paper describes the conceptual design of a production process in which waste cooking oil is
converted via supercritical transesterification with methanol to methyl esters (biodiesel).
Since waste cooking oil contains water and free fatty acids, supercritical transesterification offers
great advantage to eliminate the pre-treatment capital and operating cost.
A supercritical transesterification process for biodiesel continuous production from waste cooking
oil has been studied for three plant capacities (125,000; 80,000 and 8000 tonnes biodiesel/year). It can
be concluded that biodiesel by supercritical transesterification can be scaled up resulting high purity
of methyl esters (99.8%) and almost pure glycerol (96.4%) attained as by-product.
The economic assessment of the biodiesel plant shows that biodiesel can be sold at US$ 0.17/l
(125,000 tonnes/year), US$ 0.24/l (80,000 tonnes/year) and US$ 0.52/l for the smallest capacity
(8000 tonnes/year).
The sensitive key factors for the economic feasibility of the plant are: raw material price, plant
capacity, glycerol price and capital cost.
Overall conclusion is that the process can compete with the existing alkali and acid catalyzed
processes.

Corresponding author. Tel.: +31 13 4662203; fax: +31 13 466 3499.


E-mail address: j.m.n.v.kasteren@tue.nl (J.M.N. van Kasteren).

0921-3449/$ see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.resconrec.2006.07.005
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 443

Especially for the conversion of waste cooking oil to biodiesel, the supercritical process is an
interesting technical and economical alternative.
2006 Elsevier B.V. All rights reserved.

Keywords: Biodiesel; Waste cooking oil; Supercritical transesterification; Process design

1. Background

It is estimated that in the coming years, the fossil oil price will increase because the oil
production cannot meet the projected demand due to oil depletion (Association of Peak
Oil and Gas, 2004). This is a result of overconsumption in the developed countries and
overpopulation in the developing countries (Korbitz, 1999).
A lot of efforts have been carried out to develop an alternative fuel for the current energy
and transportation vehicle system, i.e.: fuel cell, electric power, hydrogen or natural gas for
internal combustion engines, etc. One of the promising alternatives that are applied in small
scale production is biodiesel.
The American Society for Testing and Materials (ASTM) defines biodiesel fuel as
monoalkyl esters of long chain fatty acids derived from renewable lipid feed stocks, such as
vegetable oil or animal fat. Bio represents its renewable and biological source in contrast
to traditional petroleum based diesel fuel; diesel refers to its use in diesel engines. As an
alternative fuel, biodiesel can be used in neat form or mixed with petroleum based diesel.
Several sources for producing biodiesel have been studied such as rape seed, coal seed,
palm oil, sunflower oil, waste cooking oil, soybean oil, etc. Due to the high cost of the fresh
vegetable oil, waste cooking oil gives interesting properties because it can be converted to
biodiesel and it is available with relatively cheap price (Nisworo, 2005; Zhang et al., 2003).
The most common way to produce biodiesel is by transesterification, which refers to a
catalyzed chemical reaction involving vegetable oil and an alcohol to yield fatty acid alkyl
esters (biodiesel) and glycerol (by-product) as can be seen in Fig. 1.

Fig. 1. Transesterification reaction of triglyceride and methanol to fatty acid methyl esters (biodiesel) and
glycerol.
444 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

Transesterification reactions can be alkali-catalyzed, acid-catalyzed or enzyme catalyzed


(Bunyakiat et al., 2006). An excess of methanol is used to shift the reaction to the right side
in order to achieve high yield of methyl esters/biodiesel.
Most biodiesel industries use the alkali catalyzed process. One limitation to the alkali
catalyzed process is its sensitivity to both water and free fatty acids. Free fatty acids can
react with the alkali catalyst to produce soaps and water.
Therefore Freedman et al. (1984) stated that refined vegetable oils with free fatty acids
content of less than 0.5% (acid value less than 1) should be used to maximize methyl esters
formation.
The presence of water may cause ester saponification and can consume the catalyst and
reduce the catalyst efficiency. The presence of water has a greater negative effect than that
of the free fatty acids. Ma et al. (1998) stated that the water content should be kept below
0.06%.
Most industries use pre-treatment step to reduce the free fatty acid and water content of the
feed stream. Usually free fatty acid is reduced via an esterification reaction with methanol in
the presence of sulfuric acid. The pre-treatment step not only causes the production process
to be less efficient (Kusdiana and Saka, 2004) but also increase the capital cost.
These facts hinder the efficient use of waste cooking oil, animal fats and crude oils as
source for biodiesel since they generally contain water and free fatty acids.
There is an alternative for biodiesel production, namely the supercritical methanol
method. The great advantages of supercritical methanol are:

- no catalyst required;
- not sensitive to both water and free fatty acid;
- free fatty acids in the oil are esterified simultaneously.

A comparison of the properties of the supercritical and conventional method can be seen
in Table 1.
The absence of pre-treatment step, soap removal, and catalyst removal can significantly
reduce the capital cost of a biodiesel plant, but the expected high operating cost due to
high temperature and pressure can be a drawback for supercritical method. That is why it
is interesting to see whether the supercritical methanol method is economically feasible to
be applied in a biodiesel plant.

Table 1
Properties of supercritical and conventional transesterification
Properties Supercritical Conventional
Catalyst need No (+) Yes
Reaction time Secondsminutes Minuteshours
Temperature ( C) 200300 5080
Pressure (bar) 100200 1
Free fatty acid sensitive No (+) Yes
Water sensitive No (+) Yes
Pre-treatment No (+) Yes
Catalyst removal No (+) Yes
Soap removal No (+) Yes
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 445

2. Design assumptions

Production capacities are 125,000; 80,000 and 8000 tonnes methyl esters/year.
8000 working hours per year was used.
Pressure drop of the process equipments is neglected.
Water contamination was neglected as waste cooking oil contains low water content and
the process is not water sensitive.
Assumed waste cooking oil density is 953 kg/m3 and methyl esters (biodiesel) density
of 840 kg/m3 .
Universal quasi-chemical (UNIQUAC) thermodynamic model was used due to the pres-
ence of highly polar components, i.e.: methanol and glycerol.
The feed assumed to be 100% solid particle free.
Assumed constant input of the waste oil feed for the whole production year.

3. Conceptual process design

The mass balance of the biodiesel production process is shown in Fig. 2.


The plant capacity was chosen based on the availability of waste cooking oil in The
Netherlands (cbs; Dutch Central Bureau of Statistics, 2005). Based on the paid tax of waste
cooking oil and animal fats value from the year 1999 to 2001 (20 Euro cent tax/l), a volume
of 500 million liter of waste cooking oil and animal fat was calculated for The Nether-
lands. With the density of 953 kg/m3 , the available waste cooking oil in The Netherlands is
457,440 tonnes/year.
Some of the waste cooking oil and animal fat is recycled and used as a fertilizer, soap,
and filler for cosmetics industry. That is why design capacities of 125,000; 80,000 and
8000 tonnes/year are considered realistic for the process design and simulation. The costs
of these capacities will be compared with the values of conventional biodiesel plant and
will be studied to determine the effect of capacities on the economic feasibility of the plant.
Complete process simulation was carried out to assess the commercial feasibilities of
the proposed processes.
The process simulation software, Aspen Plus Version 11.1.1 developed by Aspen Tech-
nology Inc., Cambridge, Massachusetts, USA, was used in this research.
The procedures for process simulation mainly involve defining the chemical components,
selecting a thermodynamic model, determining plant capacity, choosing proper operating
units and setting up input conditions (flow rate, temperature, pressure and other conditions).
Information on most components, such as methanol, glycerol, propane, and water is
available in the Aspen Plus component library.
Regarding the waste cooking oil or animal fat feedstock, oleic acid is considered as the
major component of the oils and fats used in the food industry in The Netherlands and Europe
(Nisworo, 2005). Triolein (C57 H104 O6 ) was chosen to represent the waste cooking oil or
animal fat in the Aspen Plus simulation. Methyl oleate (C19 H36 O2 ) was chosen to represent
the fatty acid methyl ester (biodiesel) product (as in agreement with Zhang et al., 2003).
Fig. 3 shows the process flow diagram (PFD) of the production plant. First, waste cooking
oil is preheated in a heat exchanger (B17) to 40 C to decrease the viscosity and improve
446 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

Fig. 2. Mass balance of supercritical transesterification process for waste cooking oil conversion to biodiesel, with
a yearly capacity of 125,000 tonnes biodiesel/year.

the flow property. Oil methanol mol ratio used in this process design is 1:24 and propane
(propane methanol molar ratio 1:20) used as co-solvent.
Propane is chosen as a co-solvent because it was proven to decrease the supercritical
temperature from 320 C to 280 C, the supercritical pressure from 400 to 128 bar, and the
methanol to oil ratio from 42 to 24 (mol base), respectively (Cao et al., 2005). According
to the experiments (Cao et al., 2005) in pressurized autoclave, biodiesel yield was 98% in
10 min reaction time.
Fresh streams of oil and make up of methanol are pumped to 5 bar pressure and mixed in
a mixer (B9) with the recycle stream of methanol and propane from the transesterification
reaction and accommodated in the main reactor section (details can be seen in Nisworo
(2005)). Propane solvent is soluble at 40 C and 5 bar which leads to a reduction of a power
consuming compressor.
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 447

Fig. 3. Stream 1 is input fresh waste cooking oil, stream 2 is input of fresh methanol, stream 23 is output stream
of glycerol and stream 24 is output stream of biodiesel product.

A cascade of heat exchangers is used to integrate heat of the process. The transesterifi-
cation reaction is carried out in a tubular reactor (B1). Methanol and propane are recycled
using a flash evaporator (B16) and a normal distillation column (B8). Finally biodiesel and
glycerol are obtained from settler unit (B11). The operating units will be explained further
in the coming sections.

Table 2
Design parameters of the supercritical transesterification reactor
Properties Unit
Design capacity tonnes/year 125,000 80,000 8000
Temperature C 280
Pressure Bar 128
Oil:methanol ratio Molar ratio 1:24
Propane:methanol ratio Molar ratio 1:20
Tubular reactor
Tube internal diameter cm 10 10 10
Tube thickness mm 7 7 7
Tube length m 96 58 6
Number of tubes 21 21 21
Activation energy kJ/kmol 38,482
Heat of reaction (slightly endothermic) kJ/s 0.032
Reaction kinetics constant s1 7 103
Residence time min 17
448 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

4. Supercritical transesterication and reactor

Knowing the yield and kinetics of supercritical experimental results of Kusdiana and
Saka (2001), the residence time and the dimension (length, diameter) of the reactor was
calculated and presented in Table 2.
The reactor was modeled as an adiabatic plug flow reactor (RPlug) in Aspen Plus. The
transesterification reaction, kinetics constant and activation energy were entered as the
reaction input. The activation energy was obtained by the calculation of the data of the
Arrhenius plot of Kusdiana and Saka (2001) (details in Nisworo, 2005).

5. Methanol and propane separation

The high pressure of the reactor output stream (11) was decreased to 5 bar inside a flash
evaporator (B16). This pressure drop results in transfer of the liquid propane and methanol
into gaseous form, which comes out upstream from the flash evaporator (22), containing
mainly methanol (89%) and propane (11%).
A normal distillation column (B8) is needed to separate the remaining methanol and
propane. The separation is carried out at atmospheric pressure with four stages and reflux
ratio of 0.005. These operating conditions are chosen based on the sensitivity analysis
results of the operating unit; it leads to the optimum separation of methanol and propane
mixture which still delivers biodiesel product with methanol content lower than maximum
allowable by the European biodiesel standard EN 14214 (Nisworo, 2005). The upstream
13 from distillation column (B8) contains 99% methanol and 1% propane.
Stream 13 is compressed to 5 bar before it enters mixer B4 with the stream 22. The high
temperature of stream 21 is cooled immediately with the help of a heat exchanger (B6)
using cold stream 17 (the feed stream of oil, methanol and propane). The stream 20 is 94 C
so it needs to be cooled down with the help of cooling water stream 10 in a heat exchanger
(B10). Stream 4 contains liquid methanol and propane mixture at 40 C and 5 bar and then
mixed with the fresh feed of oil and make up of methanol in mixer B9.
The total recovery of the methanol and propane is 99.3%. The make up of methanol
reacts with triglyceride to produce methyl esters and glycerol.

6. Glycerol separation

Stream 14 (17.2 tonnes/h) which contains mainly biodiesel as the end-product and glyc-
erol as by-product needs to be cooled down to 25 C using 17.2 tonnes/h cooling water in
a heat exchanger B13.
The outlet of the cooling water stream is used further to cool the methanolpropane
recycle stream, the stream leaving the reactor, and to preheat the cold feed stream of waste
cooking oil. The stream is cooled down with the help of a cooling tower (B20). Then it is
pumped back to heat exchanger B13.
Glycerol is separated using a settling tank as also very often and usually applied in
the current biodiesel industries. Stream 23 contains pure glycerol (96.4%) as a by-product
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 449

Table 3
The properties of the biodiesel end-product stream
Component End stream 24 European biodiesel standard (EN 14214)
Biodiesel (mass%) 99.82 96.50
Methanol (mass%) 0.17 0.20 (maximum)
Glycerol (mass%) 0.01 0.20 (maximum)
Triglyceride (mass%) 0.00 0.20 (maximum)
Propane (mass%) 0.00

and stream 24 contains high purity biodiesel (99.8%) and passes the European biodiesel
standard EN 14214 as can be seen in Table 3.

7. Heat balance

To conclude the process design aspect, the energy requirements for the main processing
units are described in Table 4.
The design of the process equipments are described in details in Nisworo (2005).
High reboiler duty is supplied by burning the biodiesel (assumed thermal efficiency 40%
and well-known calorific value of 37 MJ/kg). A 4 wt.% of biodiesel produced used to supply
the reboiler duty of the distillation column (Nisworo, 2005).

8. Economic evaluation

Variables in the cost calculation are: plant location, production capacity, bio-ethanol
scenario. The plant locations were simulated in the average United States location and in
The Netherlands.
These variables were studied for the sensitivity analysis. It is interesting to know the
feasibility of using bio-ethanol instead methanol (which derived from natural gas, emits
fossil CO(2) in the process).
Warabi et al. (2004) studied the reactivity of triglycerides and fatty acids of rapeseed oil in
supercritical alcohols. Using supercritical ethanol, 98% yield of ethyl esters was obtained in
45 min instead of 15 min using supercritical methanol (at their lab experiments conditions).
Assuming that the process condition is similar to the ones of methanol, only residence
time is longer (factor of 3), the bio-ethanol cost calculation was performed using multipli-
cation of factor 3 of the purchased reactor cost.

8.1. The capital cost calculation

The equipments cost which contributes to the capital cost was calculated from the data of
DACE price book (DACE (Dutch Association of Cost Engineers)), edition November 2003.
The cost was corrected with the CEPCI ratio (CEPCI stands for Chemical Engineerings
Plant Cost Index) as can be seen in Table 5.
450
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458
Table 4
Energy requirements of operating units (Nisworo, 2005)
Block name Description 125,000 (tonnes biodiesel/year) 80,000 (tonnes biodiesel/year) 8000 (tonnes biodiesel/year)

Input Out Input Out Input Out

kWth kWe kWth kWth kWe kWth kWth kWe kWth


B1 Reactor 0 0 0 0 0 0
B2 Pump 327 215 25
B8 Distillation column 2483 reboiler 33 Condenser 1609 24 182 21
B11 Settler 43 32 25
B14 Compressor 655 429 52
B16 Flash 0 0 0 0 0 0
B19 Pump 10 8 4
B20 Cooling tower 1900 1242 148
Total 2483 992 1976 1609 652 1298 182 81 194
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 451

Table 5
Chemical Engineerings Plant Cost Index
Year CEPCI
1998 389.5
2003 402.0
2004 444.2
March 2005 468.3
Source: Chemical Engineering. Economic Indicators (July 2005).

The compressor cost which is not included in the DACE price book were estimated
from the price data of process design course material given in Eindhoven University of
Technology, The Netherlands (Dautzenberg, 2003).
Table 6 shows the purchase costs of operating units and the fixed capital of the plants.
Fixed capital for equipment cost or inside battery limit (ISBL) is the cost for process-
ing units, i.e. reactors, mixers, heat exchangers, pumps, compressors, etc. This cost was
calculated by multiplying the purchased total equipment cost with a factor of 5, so

Fixed capital for equipment cost or ISBL = purchased cost 5

This factor is in agreement with 4.7 Lang factor for fluids processing plant (Sinnot,
1998).

Table 6
Fixed capital of the biodiesel plants
Equipments Code Plant capacity (tonnes/year)

125,000 125,000 80,000 80,000 8000 8000


methanol bio-ethanol methanol bio-ethanol methanol bio-ethanol
Reactor B1 97,884 293,652 74,889 224,668 18,811 56,434
Flash evaporator B16 78,074 78,074 59,733 59,733 15,004 15,004
Distillation column B8 36,763 36,763 28,127 28,127 7,065 7,065
Settler tank B11 53,102 53,102 40,627 40,627 10,205 10,205
HeatX 1 B6 398,204 398,204 304,659 304,659 76,527 76,527
HeatX 2 B10 59,445 59,445 45,481 45,481 11,424 11,424
HeatX 3 B15 54,537 54,537 41,725 41,725 10,481 10,481
HeatX 4 B17 34,445 34,445 26,353 26,353 6,620 6,620
Cooling tower B20 118,272 118,272 90,488 90,488 22,730 22,730
Compressor B14 709,527 709,527 542,846 542,846 136,357 136,357
Pump 1 B3 58,842 58,842 45,019 45,019 11,308 11,308
Pump 2 B5 13,060 13,060 9,992 9,992 2,510 2,510
Pump 3 B18 6,785 6,785 5,191 5,191 1,304 1,304
Pump 4 B12 6,785 6,785 5,191 5,191 1,304 1,304
Pump 5 B19 6,785 6,785 5,191 5,191 1,304 1,304
Equipments cost 1,732,510 1,928,278 1,325,512 1,475,291 332,954 370,576
ISBL 8,662,549 9,641,389 6,627,560 7,376,453 1,664,768 1,852,881
OSBL 1,732,510 1,928,278 1,325,512 1,475,291 332,954 370,576
Fixed capital 10,395,058 11,569,666 7,953,072 8,851,744 1,997,721 2,223,457
The values are given in US$.
452 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

The cost calculation method also includes Outside Battery Limit (OSBL) which covers
tankage, yards, roads, and other general facilities. The normal default value is 20% of ISBL.

The fixed capital = ISBL + OSBL

8.2. Operating cost calculation


The total plant capital cost = fixed capital + working capital + start-up cost

Working capital is the fund required for routine operation, including inventories, accounts
receivable and payable and cash on hand.
The result of the calculation can be seen in Table 7 .
S, G & A stands for sales, general and administrative cost which represents expenses,
research and development (R&D), administrative cost beyond plant level and corporate
overhead. It is 5% of the selling price (total operating cost and capital charges). Further
details can be found in Nisworo (2005).
The required selling price (RSP) is the price of the product which is required to cover all
costs (variable, fixed and overhead), recover the total investment and provide the specified
return of the employed capital. It is also often called break even price.
Assumed that the density of the biodiesel (methyl esters) is 840 kg/m3 , the RSP in 15 years
project life for plant capacity of 125,000 tonnes/year is US$ 0.17/l, for 80,000 tonnes/year
is US$ 0.24/l, and US$ 0.52/l for 8000 tonnes/year. Using bio-ethanol for the reaction
input, the prices are slightly more expensive, US$ 0.18/l (125,000 tonnes/year), US$ 0.25/l
(80,000 tonnes/year), and US$ 0.54/l (8000 tonnes/year).
With the same cost calculation method, the required selling price of biodiesel in The
Netherlands is summarized in Fig. 4.

Fig. 4. Required selling price of biodiesel via supercritical transesterification.


Table 7
The required selling price of biodiesel by supercritical transesterification in US

J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458
Plant capac- 125,000 tonnes/ 125,000 tonnes/ 80,000 tonnes/ 80,000 tonnes/ 8000 tonnes/ 8000 tonnes/
ity/process year methanol year bio-ethanol year methanol year bio-ethanol year methanol year bio-ethanol
condition
Fixed capital 10,395,058 11,569,666 7,953,072 8,851,744 1,997,721 2,223,457
Working capital 1,661,348 1,743,555 1,513,014 1,567,628 313,729 319,398
Start up cost 4,984,045 5,230,664 4,539,042 4,702,884 941,187 958,193
Total capital cost 17,040,452 18,543,885 14,005,128 15,122,255 3,252,638 3,501,048
Location United States
Annual variable cost
Raw material
Waste cooking oil 26,068,993 26,068,993 16,784,050 16,784,050 1,709,129 1,709,129
Methanol 4,218,750 2,736,000 278,400
Bio-ethanol 4,990,494 3,236,502 329,328
Total raw material cost 30,287,743 31,059,487 19,520,050 20,020,552 1,987,529 2,038,457
Start up
Methanol/bio-ethanol 14,400 17,040 9,050 10,710 924 1,093
Propane 4,409 4,409 2,672 2,672 269 269
Total start up cost 18,809 21,449 11,722 13,382 1,193 1,362
Utilities
Electricity 713,592 713,592 456,699 456,699 45,670 45,670
Cooling water 102,708 102,708 67,103 67,103 8,217 8,217
Biodiesel for the reboiler 978,359 1,034,908 872,157 911,252 192,657 198,371
Total utilities cost 1,794,660 1,851,209 1,395,958 1,435,054 246,544 252,257
By-product credit
Glycerol 15,937,500 15,937,500 6,234,000 6,234,000 1,017,600 1,017,600
Total by-product credit 15,937,500 15,937,500 6,234,000 6,234,000 1,017,600 1,017,600

453
454
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458
Fixed cost
Operating labor 1,020,000 1,020,000 1,020,000 1,020,000 1,020,000 1,020,000
Maintenance 485,103 539,918 371,143 413,081 371,143 371,143
Plant overhead 913,021 923,984 890,229 898,616 890,229 890,229
Taxes and insurance 207,901 231,393 159,061 177,035 39,954 44,469
Total fixed cost 2,626,024 2,715,295 2,440,433 2,508,733 2,321,326 2,325,841
Total operating cost 18,789,736 19,709,939 17,134,165 17,743,720 3,538,992 3,600,318
Capital chargesa 5,314,381 5,787,437 4,353,483 4,707,127 1,207,569 1,287,014
S, G and A 1,205,206 1,274,869 1,074,382 1,122,542 237,328 244,367
Required selling price 25,309,323 26,772,245 22,562,030 23,573,389 4,983,890 5,131,699
RSP (US$/tonnes) 202 214 282 295 623 641
RSP (US$/kg) 0.20 0.21 0.28 0.29 0.62 0.64
RSP (US$/l) 0.17 0.18 0.24 0.25 0.52 0.54
a 20% return of investment (ROI) used.
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 455

From Fig. 4, it can be concluded that the biodiesel produced via supercritical transester-
ification which use no catalyst and no pre-treatment step is economically feasible since the
biodiesel can be sold at 17 US cent/l for the 125,000 tonnes/year capacity.
Biodiesel produced from small plant (8000 tonnes/year) in The Netherlands is the most
expensive compared with the others.
For all capacities, biodiesel price in The Netherlands is more expensive than in the United
States. This can be attributed to the high price of waste cooking oil; 30 Euro cent/l (Rice et
al., 1998) which is equal to 37 US cent/l. The price of the waste cooking oil in the United
States is 20 US cent/l (Zhang et al., 2003).
Bio-ethanol hardly influences the biodiesel price as it only 1 US$ cent more expensive per
liter. This can be attributed to considerably small sensitivity of the methanol or bio-ethanol
as the reactants ranging between 7 and 14% (Nisworo, 2005) although the bio-ethanol price
is more expensive compared with methanol (bio-ethanol price US$ 355/tonnes, methanol
price US$ 300/tonnes).
From Fig. 5, it can be seen that the biodiesel produced via non-catalytic way (49 US
cent/l) can compete with the industrial way produced biodiesel (Zhang 01 and 02 bars)
(Zhang et al., 2003a) as it is cheaper to sell. The method most used in biodiesel industries
is alkali catalyst with acid pre-treatment step prior to main reaction.
The high cost of the biodiesel (72 US cent/l, second bar from the left) can be attributed
to the high cost of the fresh vegetable oil.
Whereas the required selling price of 74 US cent/l is due to the catalyst cost and additional
pre-treatment step to reduce the free fatty acid of the waste cooking oil although the cost of
the waste cooking oil is lower than fresh vegetable oil.

Fig. 5. The required selling prices of small biodiesel plants.


456 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

Table 8
Summary of the sensitive key factors as percentage of manufacturing cost
tonnes/year The Netherlands United States Bio-ethanol
125,000 Waste oil (80%) Waste oil (73%), Waste oil
capital (15%) (71%), capital
(16%)
80,000 Waste oil (77%) Waste oil (68%), Waste oil
capital (16%) (67%), capital
(17%)
8000 Waste oil (49%), capital (18%), Waste oil (35%), Waste oil
human labor (16%) capital (20%), human (34%), capital
labor (21%) (21%), human
labor (21%)

Table 9
Percentage reduction of operating cost by by-product sale
tonnes/year The Netherlands United States Bio-ethanol
125,000 Glycerol (22%) Glycerol (36%) Glycerol (34%)
80,000 Glycerol (13%) Glycerol (20%) Glycerol (20%)
8000 Glycerol (12%) Glycerol (16%) Glycerol (15%)

Zhang et al. (2003) also carried out process design and economic feasibility with acid
catalyst process which requires longer reaction time from waste cooking oil feed. It is
insensitive to low free fatty acids in the waste cooking oil.
From Fig. 5, it is clear that biodiesel produced non-catalytically is 5 US cent cheaper
than biodiesel produced with acid catalyst process (Zhang 03 bar).
Kusdiana and Saka (2001) experiment with acid catalyst shows that waste cooking
oil which contains 10% of free fatty acids resulted in biodiesel yield drop to 71%.
Whereas their supercritical experiment with the same waste cooking oil resulted in 98%
yield.
Zhang et al. (2003) based their calculation on 97% yield which seems not realistic, so in
reality the price of the acid catalyzed biodiesel will be much higher.
The major cost contributor to the price of biodiesel can be seen Table 8.
For big plants, the major cost contribution for the biodiesel price is the price of the raw
material (7180%) and capital charges (1516%). Glycerol as the by-product also influences
the price of the biodiesel as it contributes 2236% reduction of the biodiesel price (Table 9).
For small plant capacity, the contribution of operating labor, maintenance, overhead
become higher compared with bigger capacities.
This means that plant capacity changes the sensitive factors for cost contribution.

9. Conclusions

A supercritical transesterification process for biodiesel continuous production from waste


cooking oil has been studied for three plant capacities (125,000; 80,000 and 8000 tonnes
J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458 457

biodiesel/year). It can be concluded that biodiesel by supercritical transesterification can be


scaled up resulting high purity of methyl esters (99.8%) and almost pure glycerol (96.4%)
attained as by-product.
The economic assessment of the biodiesel plant shows that biodiesel can be sold at
US$ 0.17/l (125,000 tonnes/year), US$ 0.24/l (80,000 tonnes/year) and US$ 0.52/l for the
smallest capacity (8000 tonnes/year).
The sensitive key factors for the economic feasibility of the plant are: raw material price,
plant capacity, glycerol price and capital cost.
Overall conclusion is that the process can compete with the existing alkali and acid
catalyzed processes.
Especially for the conversion of waste cooking oil to biodiesel, the supercritical process
is an interesting technical and economical alternative.

Acknowledgements

The authors are much indebted to Johan Wijers for the help with Aspen simulation
software.

References

Association of Peak Oil and Gas (ASPO); updated 2004 (Internet source: http://www.peakoil.net/uhdsg/
Default.htm).
Bunyakiat K, Makmee S, Sawangkeaw R, Ngamprasertsith S. Continuous production of biodiesel via transesteri-
fication from vegetable oils in supercritical methanol. Energy Fuels 2006;20:8127.
Cao W, Han H, Zhang J. Preparation of biodiesel from soybean oil using supercritical methanol and co-solvent.
Fuel 2005;84(2005):34751.
Chemical Engineering. Economic Indicators, 2005;112(July(7)):634 (www.che.com).
Dautzenberg FM, 2003. Process Design and Scale Up Course Material. New Jersey: ABB Lummus (Copyright of
Eindhoven University of Technology [printed in the Netherlands]).
Dutch Central Bureau of Statistics Internet Source. the Netherlands; 2005 (www.cbs.nl).
Freedman B, Pryde EH, Mounts TL. Variables affecting the yields of fatty esters from transesterified vegetable
oils. J Am Oil Chem Soc 1984;61:163843.
Korbitz W. Biodiesel production in Europe and North America, an encouraging prospect. Renew Energy
1999;16:107883.
Kusdiana D, Saka S. Effects of water on biodiesel fuel production by supercritical methanol treatment. Bioresour
Technol 2004;91(2004):28995.
Kusdiana D, Saka S. Methyl Esterification of free fatty acids of rapeseed oil as treated in supercritical methanol.
J Chem Eng Jpn 2001;34(3):3837.
Ma F, Clements LD, Hanna MA. The effects of catalyst, free fatty acids, and water on transesterification of beef
tallow. Trans ASAE 1998;41:12614.
NAP DACE Prijzenboekje. 23rd ed.; November 2003. the Netherlands: DACE (Dutch Association of Cost Engi-
neers).
Nisworo AP, 2005. Biodiesel by supercritical transesterication: process design and economic feasibility. Twaio
graduation project thesis report. Eindhoven University of Technology.
Rice B, Frohlich A, Leonard A. Bio-diesel production from camelina oil, waste cooking oil and tallow. Oak Park,
Carlow: Crops Research Centre; 1998.
Sinnot RK, 1998. Coulson & Richardsons Chemical Engineering, vol. 6.
458 J.M.N. van Kasteren, A.P. Nisworo / Resources, Conservation and Recycling 50 (2007) 442458

Warabi Y, Kusdiana D, Saka S. Reactivity of triglycerides and fatty acids of rapeseed oil in supercritical alcohols.
Bioresour Technol 2004;91(2004):2837.
Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste cooking oil. 1. Process design and
technological assessment. Bioresour Technol 2003;89(2003):116.
Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste cooking oil. 2. Economic assessment
and sensitivity analysis. Bioresour Technol 2003a;90(2003):22940.

You might also like